Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Ocean Engineering 235 (2021) 109439

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Numerical investigation on nonlinear wave interaction with partially


submerged subsea pipe-valve structures
Yang Du a, *, Miaozi Zheng b, Bingqi Liu c, Carlos Levi c, HyoJae Jo d, Menglan Duan a
a
College of Safety and Ocean Engineering, China University of Petroleum-Beijing, Beijing, China
b
Engineering & Design Institute of CNPC Offshore Engineering Company Limited, Beijing, China
c
Ocean Engineering Department, COPPE, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil
d
Department of Naval Architecture and Ocean Systems Engineering, Korea Maritime and Ocean University, Busan, South Korea

A R T I C L E I N F O A B S T R A C T

Keywords: The dynamic behaviours of subsea pipe-valve structures—the principal components of subsea equipment—are of
Subsea pipe-valve structures critical concern for the control of the lowering through the splash zone during installation. Nonlinear interactions
Partially submerged between waves and pipe-valve combinations are challenging the investigation of the wave forces acting on large-
Numerical water tank
dimension structures. This paper proposes a numerical model to study the nonlinear hydrodynamic performance
Wave-structure interaction
Nonlinearity
of partially submerged pipe-valve structures. A two-dimensional (2D) numerical water tank (NWT) was estab­
lished to efficiently generate high-accuracy incident waves, and a semi-immersed horizontal cylinder was applied
to the numerical simulations of wave-structure interaction (WSI) to study the mesh and time convergence. The
numerical model was validated with a comparative analysis of the modelling results with experimental data,
other numerical results, and theoretical results. Numerical simulations of the interaction between Stokes-II waves
and subsea pipe-valve structures were performed, and the effects of the valve height ratio, valve length ratio, and
valve spacing were investigated. The subsea pipe-valve structure is a kind of multi-scale structure, the wave
forces exhibited stronger nonlinearity when acting on the partially submerged valves than on the completely
submerged pipe. Changes in valve structure and distribution significantly increased the overall change in the
wave forces acting on the valves. The stronger nonlinearity was primarily derived from phenomena such as wave
reflection, wave blockage, and wave run-up. The numerical results provide references for the structure and
strength designs and offshore installations of large subsea equipment like subsea manifolds.

1. Introduction approaching the seabed (Veritas, 2011). The nonlinear wave force
acting on subsea structures in the second phase may result in structural
The development of offshore oil industry technology and the damage, such as deformation and bending. Therefore, more information
increased water depth of offshore oilfields have accelerated the devel­ on WSI must be provided before conducting practical offshore
opment of subsea production systems to exploit deepwater oil and gas installations.
resources. Subsea production systems play a significant role in the WSI is a typical problem in ocean and coastal engineering that has
extraction of deepwater hydrocarbons, have a wide application range been studied by many researchers using different approaches, such as
and high efficiency, and are less affected by environmental loads (Cao analytical solutions, physical experiments, and numerical simulations.
et al., 2016; Hong et al., 2018; Lindseth et al., 2019; Jamaludin, 2019). Buccino and Martinelli (2020) summarised the arguments of the papers
Subsea production systems are composed of vital subsea equipment, gathered in a special issue named interaction between waves and
such as subsea manifolds, Xmas trees, and pipeline end terminations maritime structures, primarily the structural and functional perfor­
(Mason and Upchurch, 1996; Bernt, 2004). The four primary phases of mances; additionally, regarding the functional performance, it was
lowering subsea equipment from the installation ship to the seabed subdivided into the hydraulic, morphodynamic, and floating body re­
consist of lifting equipment from the deck of the installation ship, sponses. Analytical method is relatively more cost-effective and efficient
lowering through the splash zone, lowering in deep waters, and for studying WSI under certain conditions and provides a good scientific

* Corresponding author.
E-mail address: 2017310001@student.cup.edu.cn (Y. Du).

https://doi.org/10.1016/j.oceaneng.2021.109439
Received 18 February 2021; Received in revised form 31 May 2021; Accepted 29 June 2021
Available online 6 July 2021
0029-8018/© 2021 Elsevier Ltd. All rights reserved.
Y. Du et al. Ocean Engineering 235 (2021) 109439

understanding of the hydrodynamic characteristics of structures in the comparing the differences in absorption quality between these methods.
WSI process. Fang et al. (2018) presented an analytical solution based on Aggarwal et al. (2019) used the REEF3D to conduct CFD-based numer­
the linear potential theory to predict hydrodynamic wave forces sub­ ical simulations to study breaking wave characteristics and the influence
jected by girder-type bridge decks under hurricane-generated oblique of the breaker location on the wave impact loads of a jacket structure.
water waves. Jalón et al. (2019) considered the linear wave theory and The numerical simulations in this study were performed using Open­
the head loss caused by the constriction of flow to develop an analytical FOAM, a free and open-source CFD package that uses the finite volume
model of the oblique incident wave interaction with a maritime struc­ method to solve the discretised Navier–Stokes equations. Extensive
ture and extended it to irregular waves. Bahena-Jimenez et al. (2020) numerical simulations involving the interaction between waves and
obtained formulas that obey a power-law function for the reflection and structures have been reported based on the OpenFOAM software pack­
transmission coefficients and the wave elevation for the linear water age. Hu et al. (2016) used OpenFOAM to simulate WSIs and developed a
wave interaction with a cycloidal breakwater in presence of a sloping new boundary condition for extreme waves. The new boundary condi­
beach. However, the analytical method has limitations, such as the po­ tion was implemented to simulate the interaction with a fixed/floating
tential flow models established under the assumption of inviscid and truncated cylinder and a simplified FPSO. Islam et al. (2019) used
irrotational fluid flow, and cannot adequately address the velocity fields OpenFOAM to analyse wave radiation due to heave, surge, and pitch by
around structures and the strong nonlinear effect. Li et al. (2020a) a box-type floating structure based on CFD simulations and applied the
analytically and experimentally investigated the interaction between waves2Foam toolbox to generate and absorb free surface waves. Based
water waves and a submerged perforated quarter-circular caisson on OpenFOAM package, Gao et al. (2020) used a 2D numerical wave
breakwater. Wave energy dissipation near the breakwater crown is flume to numerically investigate wave loads during gap resonance be­
considerable owing to flow separation but cannot be considered in the tween a fixed box and a vertical wall, and employed the waves2Foam
potential solution. Therefore, at a water depth of 0.5 m, the reflection toolbox to produce incident waves at the boundaries and avoid wave
and transmission coefficients calculated by the analytical solution are re-reflection in the water flume.
slightly larger than those measured by the experimental tests. The operational environment during the lowering and installation of
Extensive studies on physical WSI experiments have been reported. subsea equipment is often harsh. Especially, strong nonlinear phenom­
Based on physical water tanks, different wave generators such as piston, ena, such as wave breaking and splashing affect the process of lowering
flap, and plunger are used to generate different types of waves. Lamberti through the splash zone. Environmental loads can cause structural
et al. (2011) focused on wave-induced horizontal forces on the deck damage to subsea equipment, leading to installation failures or engi­
front, described a series of large-scale experiments on exposed jetties, neering accidents. In subsea production systems, subsea manifolds as the
and provided guidelines on the optimal logging frequency or low-pass critical equipment are used to simplify the system, to minimise the usage
filter to be used if the experimental loads were obtained by pressure of subsea pipelines and risers, and to optimise fluid flow during pro­
integration. Rodríguez and Spinneken (2016) conducted a series of duction. The subsea manifold is an arrangement of pipes and valves
experimental tests in regular and irregular wave conditions concerning aimed at combining, distributing, controlling, and monitoring fluid flow
the nonlinear loading and dynamic response of a heaving rectangular (Bai and Bai, 2019). The pipe-valve structure is a multi-scale critical
box in two dimensions, applied four absorbing flap-type wavemakers to component installed inside the subsea equipment, and it is necessary to
generate incident wave conditions, and obtained all experimental data investigate its hydrodynamic performance under nonlinear incident
in the long wave flume. Sayeed et al. (2017) conducted experiments to waves. For example, the subsea manifold contains different types of
study the change in wave loads for different sizes of small spherical valves (e.g., hydraulic ball valve, hydraulic gate valve, and manual ball
bodies at different proximities to a fixed structure, a hydraulically valve) of different heights and lengths. The close arrangement of the
actuated wave board and a wave absorbing beach were respectively valves causes the hydrodynamic behaviour of a valve will be affected by
equipped on both sides of the towing tank. Huang et al. (2018) experi­ adjacent valves.
mentally and numerically investigated hurricane-induced wave forces This study performed numerical calculations to investigate the
on a box girder deck of a coastal bridge. In the experiments, the periodic interaction between nonlinear waves and partially submerged pipe-
waves were produced on the left side of the wave flume using a wave valve structures. We investigated the effects of the valve height ratio,
generator without reflection. Yan et al. (2019) used a piston-type valve length ratio, and valve spacing on wave forces subjected by subsea
wavemaker to produce the plunging wave based on the focused wave pipe-valve structures in a 2D NWT and analysed the reasons for wave
theory, and experimentally and numerically studied the plunging wave force changes. The obtained results provide references for the structure
impacts on a box-shape structure. Physical experiments, a traditional and strength designs and valve layout of subsea pipe-valve structures.
research approach, are highly accurate and reliable but have limitations, The rest of this paper proceeds as follows: Section 2 presents the nu­
such as long research time, high cost, and scale restriction. merical method used in the current study, introduces the governing
The rapid development of computational fluid dynamics (CFD) and equations, uses the static-boundary method to generate Stokes-II waves,
improved computational power allow numerical simulations to provide and determines the mesh generation scheme. A series of numerical
an economic, highly efficient, and powerful alternative to perform the simulations for the interaction between waves and a semi-immersed
WSI investigations. Numerical study of WSI requires the development of cylinder are performed to study the mesh and time convergence and
a NWT, which can overcome limitations of physical model tests, such as to validate the numerical model. In Section 3, a series of numerical
model size and viscous similarity. Many researchers, including Anbar­ simulations are conducted to calculate the horizontal and vertical forces
sooz et al. (2013), Dao et al. (2018), Wu and Hsiao (2018), and Marques acting on partially submerged subsea pipe-valve structures at different
Machado et al. (2018) have conducted studies on NWTs. Finnegan and valve structures and valve distributions. Section 4 provides conclusions
Goggins (2015) used the commercial software ANSYS CFX to develop a based on the results.
CFD NWT model and introduced a rectangular prism structure into the
model to investigate a linear irregular ocean wave interaction with a 2. Numerical model
structure. Ghezelbashan and D’Mello (2017) used CFD package ANSYS
Fluent with a Stokes-V wave to model the nonlinear high-amplitude 2.1. Governing equations
ocean wave and used ANSYS Workbench to investigate wave interac­
tion with a vertical slender cylindrical member. Miquel et al. (2018) With respect to an incompressible two-phase (water and air) flow,
analysed the performance of different wave generation and absorption the densities of both fluids are constants, the governing equations are
methods in CFD-based NWTs using the open-source hydrodynamic the continuity and momentum equations. The continuity equation is
model REEF3D, and conducted a sensitivity analysis for quantifying and given by:

2
Y. Du et al. Ocean Engineering 235 (2021) 109439

∇⋅U = 0 (1) πH sinh(kd + kz) 3 π 2 H 2 sinh(2kd + 2kz)


Uv = sin(kx − ωt) + sin(2kx − 2ωt)
T sinh(kd) 4 TLsinh4 (kd)
where U is the velocity vector. In the volume of fluid method,
α(0 ≤ α ≤ 1) represents the water volume fraction in each cell. For α = 1, (8)
it means that the cell is full of water; for α = 0, it means that the cell is
where Uh and Uv represent the components of the particle velocity in the
full of air. In addition to these two cases, it represents the air-water
horizontal and vertical directions, respectively. z is the vertical coordi­
interface. Thus, Eq. (1) can be reduced to the following form:
nate used to describe wave movement, and z = 0 denotes the still water
∂α level. T is the wave period and L is the wavelength.
+ ∇ ⋅ (Uα) + ∇ ⋅ [Ur α(1 − α)] = 0 (2)
∂t Therefore, a NWT was established with a length of 20 m, width of
0.06 m, height of 1 m, and water depth of 0.6 m. Because the numerical
where Ur represents a relative velocity, and ∇⋅[Ur α(1 − α)] is an artificial simulations were conducted in 2D, the width of the NWT did not affect
compression term for limiting the interface smearing (Rusche, 2003). the simulation results. Therefore, the extrudeMeshDict and crea­
The momentum equation is given by: tePatchDict dictionaries in OpenFOAM were used to ensure that the
∂ρU computational domain had only one grid layer in the width direction to
+ ∇ ⋅ (ρUU) = ∇ ⋅ (μ∇U) − ∇p − g ⋅ x∇ρ + σ T κα ∇α (3) greatly reduce the number of grids required for the numerical calcula­
∂t
tions, thus saving the computational cost. Mesh size plays a significant
where p is the dynamic pressure; g is the gravitational acceleration; x is role in the accuracy of the numerical results; thus, mesh convergence
the position vector; σT denotes the surface tension coefficient; κα denotes was studied to determine a reasonable mesh system with high accuracy
the surface curvature of the air-water interface; and ρ and μ are the and efficiency for wave generation. Firstly, the mesh size was globally
density and viscosity of the mixture, respectively, the equations of which refined along the horizontal (x-axis) and vertical (z-axis) directions,
are as follows: respectively. The wave height was set to 0.05 m, and the wave period
was set to 3 s. The cell sizes remained at 0.01 m in the vertical direction,
ρ = αρwater + (1 − α)ρair (4)
and were set to 0.06 m, 0.04 m, 0.03 m, and 0.02 m in the horizontal
direction, respectively. Fig. 1(a) shows that mesh refinement in the
μ = αμwater + (1 − α)μair (5)
horizontal direction has little effect on wave elevation prediction ac­
curacy, and the numerical results are in good agreement with the
2.2. Wave generation based on the static-boundary method theoretical data. HD and VD represent the cell size (unit [m]) in the
horizontal and vertical directions, respectively. The simulation times
Practical wave height in the ocean is generally limited with respect to consumed by the four cell sizes (HD0.06–VD0.01, HD0.04–VD0.01,
the wavelength or water depth, and sometimes it can reach a relatively HD0.03–VD0.01, and HD0.02–VD0.01) were 647 s, 829 s, 856 s, and
large value. For the finite amplitude wave, the nonlinear effects caused 1279 s, respectively. Then, the cell sizes remained at 0.04 m in the
by the fluctuating free surface must be considered; namely, the motion horizontal direction, and were set to 0.015 m, 0.01 m, 0.007 m, and
and dynamic conditions of the free surface are nonlinear. Considering 0.005 m in the vertical direction, respectively. Fig. 1(b) shows that the
the calculational accuracy and cost, this study uses the static-boundary prediction accuracy for the wave elevation is not good under the
method to generate nonlinear waves. For wave generation and absorp­ HD0.04–VD0.015 cell size; for vertical cell sizes of less than 0.01 m, the
tion, many researchers used the waves2Foam within OpenFOAM, which numerical results change very little as the cell size decreases. The
is a third-party toolkit developed by Jacobsen et al. (2012). In this study, simulation times consumed by the four cell sizes (HD0.04–VD0.015,
we used the IHFOAM toolbox to generate and absorb incident waves, HD0.04–VD0.01, HD0.04–VD0.007, and HD0.04–VD0.005) were 482 s,
which is embedded in OpenFOAM and can be applied to generate 829 s, 1011 s, and 1290 s, respectively. The time consumed gradually
different types of waves (e.g., Cnoidal, Stokes-I, Stokes-II, and Stokes-V). increased as the mesh refinement level increased. Secondly, to reduce
Additionally, the active wave absorption of IHFOAM works simulta­ the time consumed in the simulation without reducing the calculation
neously with wave generation to dissipate incident waves. IHFOAM is accuracy, the HD0.04–VD0.01 cell size was used to capture the free
the fastest wave generation method, whereas waves2Foam is 11% surface elevation near the still water level, and the HD0.04–VD0.04 cell
slower because of the use of relaxation zones (Martínez-Ferrer et al., size was used in other domains, namely local mesh refinement. Fig. 2
2018). In the present study, combined with the setting of related pa­ compares the predicted wave elevation based on global mesh refinement
rameters and according to the range of application of the various wave (only the HD0.04–VD0.01 cell size) and local mesh refinement. The
theories proposed by Le Méhauté (1976), Stokes-II wave theory is the wave elevations predicted by the two mesh generation schemes agree
most suitable for this paper. Additionally, the Stokes-II wave theory is well, and the times consumed by the two schemes were 829 s and 445 s,
representative, widely used in the investigation of WSIs, and the basis respectively. Thus, local mesh refinement is a good mesh generation
for investigating the interactions between structures and more complex scheme that efficiently generates high-accuracy incident waves. The
waves. The corresponding free-surface elevation is expressed as follows: numerical results prove that the wave generation is mesh-convergent.
H kH 2 cosh(kd)
η= cos(kx − ωt) + [2 + cosh(2kd) ]cos(2kx − 2ωt) (6)
2 16 sinh3 (kd) 2.3. Validation

where H is the wave height, k is the wave number, x is the distance along
A numerical simulation scheme based on the above-mentioned 2D
the longitudinal direction, ω is the wave frequency; and d and t represent
NWT is presented to investigate the interaction between nonlinear
the water depth and time, respectively. Based on the static-boundary
waves and partially submerged structures. Fig. 3 shows the computa­
method, the water particle velocity was specified in the horizontal and
tional domain of the interaction between Stokes-II waves and a semi-
vertical directions on the inlet boundary for wave generation, the par­
immersed horizontal cylinder; the radius of the cylinder is 0.05 m,
ticle velocity equations based on the Stokes-II wave theory are as fol­
and the centre coincides with the still water level. Optimal boundary
lows:
conditions were applied to the numerical simulations. The left side
πH cosh(kd + kz) 3 π2 H 2 cosh(2kd + 2kz) (inlet) and right side (outlet) of the NWT were used to generate incident
Uh = cos(kx − ωt) + cos(2kx − 2ωt) waves and to avoid wave reflection, respectively; thus, the velocity
T sinh(kd) 4 TLsinh4 (kd)
(7) boundary condition was used for the inlet, and the wave mode of
shallow water absorption was used for the outlet. The wall boundary

3
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 1. Effect of global mesh refinement on wave elevation prediction accuracy.

condition was applied for the structure and the NWT bottom, and the
type keyword fixedValue was applied to ensure zero velocity at these
boundaries. The boundary condition at the top side of the NWT was set
to inletOutlet, which was used to avoid backflow. The numerical sim­
ulations in this study were performed in 2D; thus, the front and back
sides of the NWT should be set to empty. The density and kinematic
viscosity of the water were set to 1000 kg/m3 and 10− 6 m2/s, respec­
tively; and the density and kinematic viscosity of the air were set to 1 kg/
m3 and 1.48 × 10− 5 m2/s, respectively.
The mesh generation scheme proposed for wave generation may be
unsuitable for the numerical calculation of wave interaction with
structures; thus, further mesh refinement is needed around structures
where high velocity gradients may occur (Fig. 3). To study mesh
convergence, three mesh refinement levels (coarse, middle, and fine)
were applied to simulate the wave interaction with a semi-immerged
cylinder, respectively. The normalised wave force amplitude Fh,v ,
(see
Eq. (9)) was used for the mesh convergence study:
√̅̅̅
2FRMS
(9)

Fh,v =
ρwater gRAl

Fig. 2. Comparison of predicted wave elevation under different mesh refine­ where R is the cylinder radius, l is the cylinder length, A is the wave
ment schemes. amplitude. Root-mean-square (RMS) is also called the quadratic mean,
and is obtained by summing the square of all values, finding its mean
value, then calculating the root. FRMS is the RMS of the horizontal wave
√̅̅̅
force Fh or the vertical wave force Fv , and 2FRMS denotes the wave force

Fig. 3. Computational domain for wave interaction with a semi-immerged cylinder.

4
Y. Du et al. Ocean Engineering 235 (2021) 109439

amplitude. The FRMS equation can be expressed as follows: the calculations are time-convergent. Additionally, the times consumed
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ for the numerical simulations under MaxCo numbers of 1, 0.65, and 0.5
√ ∫T

√1 were 1421 s, 1714 s, and 1929 s, respectively. Due to the small
FRMS = √ F2h,v (t) (10) discrepancy in time consumed, a MaxCo number of 0.5 was selected to
T
0 guarantee stable numerical calculations.
To validate the accuracy and reliability of the numerical model,
According to the principle of Dixon et al. (1979), the theoretical
numerical simulations were conducted for the interaction between the
wave forces acting on structures in the horizontal and vertical directions
Stokes-II waves and a semi-immersed cylinder. The incident wave am­
are:
plitudes were set to 0.01 m, 0.015 m, 0.02 m, 0.03 m, 0.04 m, and 0.05
1 ∂Uh m, respectively; and the wave period was set to 1 s. In this study,
Fh = Cd ρwater Sh (t)|Uh |Uh + Cm ρwater V(t) (11)
2 ∂t considering the selection of calculation parameters, and for batter
comparing with other research results, thus the Fh,v (see Eqs. (9) and

∂Uv
Fv = Cm ρwater V(t) + ρwater g[V(t) − V0 ] (12) (10)) obtained by our present simulations (represented by P) were
∂t
compared with the linear theory results (represented by T), the experi­
where Cd and Cm are the drag and inertia coefficients, respectively. Sh (t) ment data (Martin and Dixon, 1983) (represented by E), and the pub­
is the instantaneous immersed projection area of the structure in the lished numerical results (Tan Loh et al., 2018) (represented by N)
horizontal direction. V(t) and V0 are the instantaneous and initial (Fig. 6).The difference between P and T increases as the wave amplitude
immersed volumes of the structure, respectively. The last term in Eq. increases, primarily because the theoretical results are independent of
(12) was subtracted to analyse only the dynamic force. A/R. Additionally, the linear theory is only valid for small A/R values.
The incident wave height was set to 0.05 m, and the wave period was As A/R gradually increases, the nonlinearity of the wave forces acting on
set to 1 s. Fig. 4 shows the numerical results of the horizontal and vertical the cylinder increases and cannot be accurately predicted by linear
forces acting on the cylinder under different mesh refinement levels. theory, namely the linear theory has a weaker prediction accuracy for
Regarding the prediction of Fh , calculation errors between the no mesh

large A/R values. The overall tendencies of our present results, experi­
refinement, coarse mesh, middle mesh, and fine mesh were approximately mental data, and published numerical results are similar. Meanwhile, for
18.27%, 2.32%, and 2.01%, respectively. Regarding the prediction of Fv ,
′ A/R less than 0.6, the errors between P, N, E, and T were approximately
8.12%, 14.21%, and 13.51%, respectively. Thus, our present results
calculation errors between the no mesh refinement, coarse mesh, middle
show better agreement with the theoretical data at low A/R values.
mesh, and fine mesh were approximately 8.22%, 2.48%, and 0.79%,
Furthermore, comparing P and E reveals that the minimum and average
respectively. The numerical results obtained by the middle and fine meshes
errors were approximately 0.04% and 3.04%, respectively. It is indi­
show very little discrepancy, proving that mesh convergence is guaranteed.
cated that our results agree well with the experimental results. In sum­
Additionally, the times consumed for numerical simulations under coarse,
mary, the numerical model proposed in this study can efficiently and
middle, and fine meshes were 626 s, 1714 s, and 2587 s, respectively.
accurately predict Stokes-II wave interaction with partially submerged
Considering the calculation accuracy and efficiency, the middle mesh was
structures.
selected for the following simulations.
The time step plays a significant role in the accuracy of the numerical
results; thus, the time convergence was studied to select a reasonable 3. Results and discussion
maximum courant (MaxCo) number for the numerical simulations.
Fig. 5 shows the numerical results of the horizontal and vertical forces Subsea pipe-valve structures are the critical components of subsea
acting on the cylinder under different MaxCo numbers. Regarding the equipment, and the dynamic behaviours are of critical concern for con­
prediction of Fh , calculation errors between MaxCo numbers of 1, 0.65

trolling the process of lowering through the splash zone during installation.
and 0.5 were approximately 0.71% and 0.36%, respectively. Regarding The size of valve is much smaller than the size of pipe; thus, the pipe-valve
structure is a multi-scale structure. The mechanism of the interaction be­
the prediction of Fv , calculation errors between MaxCo numbers of 1,

tween the waves and the pipe-valve structures is significantly different


0.65 and 0.5 were approximately 0.86% and 0.085%, respectively. Thus,
from that of other simple structures (e.g., cylinders and boxes). Multi-scale

Fig. 4. Mesh convergence study for wave force predictions.

5
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 5. Time step convergence study for wave force predictions.

Fig. 6. Comparison of the normalised horizontal and vertical force amplitudes acting on a cylinder.

Fig. 7. Sketch of incident waves past valve and pipe-valve structures.

6
Y. Du et al. Ocean Engineering 235 (2021) 109439

structures will produce significant multi-scale coupling phenomena when the total simulation time was set to 24 s. Additionally, the decom­
interacting with nonlinear waves. Structural interactions at different scales poseParDict dictionary in OpenFOAM was used to conduct parallel
complicate the wave motion track, produce strong nonlinear wave loads, computation, which breaks the geometry and associated fields into sub-
and increase the wave forces acting on the structures and the possibility of fields and distributes them to different processors for calculation. Par­
structural damage. allel computing primarily decomposes the grids and fields, runs the
We performed numerical calculations on the interactions between application in parallel, and post-processes the decomposed fields. All the
the incident wave and the independent valve structure and the incident simulations in this study were executed on a computer with an Intel(R)
wave and the pipe-valve structure (Fig. 7). The valve has a length of Core(TM) i7-6700HQ CPU with eight physical cores. Therefore, eight
0.06 m, a height of 0.06 m, and a centre that coincides with the still cores were used to run in parallel for every simulation case, greatly
water level. The pipe has a length of 1.5 m and height of 0.06 m; the reducing the execution time.
distance between the top of pipe and the still water level is 0.03 m. Four
wave gauges (WG1, WG2, WG3, and WG4) were deployed near the valve 3.1. Effect of valve height ratio on wave forces acting on the subsea pipe-
to record the surrounding wave fields. WG1 and WG2 were placed on the valve structure
left side of the valve; whereas WG3 and WG4 were located on the right
side. WG2 and WG3 were very close (0.02 m) to the sides of the valve. The effect of the valve height ratio on wave forces subjected by the
The incident wave height was set to 0.06 m, and the wave period was set subsea pipe-valve structure was studied by setting two valves with
to 1.5 s. Fig. 8(a) shows the calculated wave forces acting on an inde­ specified heights on a pipe (Fig. 10). LPipe and HPipe represent the length
pendent valve structure; the nonlinearity of the horizontal and vertical and height of the pipe, respectively; LValve1 and HValve1 represent the
forces acting on the valve was relatively weak. Fig. 8(b) shows that the length and height of the first valve (hereinafter referred to as valve 1),
nonlinearity of the wave forces is significantly greater on the valve of the respectively; and LValve2 and HValve2 represent the length and height of
pipe-valve structure than that on the independent valve structure, the second valve (hereinafter referred to as valve 2), respectively. D12
because of the pipe. Additionally, the maximum absolute values of the denotes the spacing between valves 1 and 2, and the numerical simu­
horizontal and vertical forces acting on the valve of the pipe-valve lations are divided into two groups. In the first group of simulations, the
structure all increase significantly, indicating that the pipe-valve struc­ height of valve 2 remained at 0.06 m, and the height of valve 1 increased
ture is more prone to structural damage than the independent valve from 0.01 m to 0.06 m at 0.01 m increments. In the second group of
structure. simulations, the height of valve 1 remained at 0.06 m, and the height of
For an independent valve structure, Fig. 9(a) shows the wave valve 2 increased from 0.01 m to 0.06 m at 0.01 m increments. The
elevation time histories at different wave gauges, it can be found that the
positive and negative signs indicate the direction of the wave forces. Fh+
nonlinearity of the wave elevations near the valve is relatively weak.
and Fv+ represent the maximum horizontal and vertical forces acting on
The recorded wave crests were larger at WG1 and WG2 than at WG3 and
the structures in a specific simulation case, respectively. Fh− and Fv−
WG4, primarily due to the wave reflection effect produced by the valve.
represent the minimum horizontal and vertical forces acting on the
Compared with the independent valve structure, Fig. 9(b) shows that the
structures in a specific simulation case, respectively. Fhmean and Fvmean
nonlinearity of wave elevations near the valve of the pipe-valve struc­
ture is significantly enhanced; additionally, the motion track of incident represent the average horizontal and vertical forces acting on the
waves becomes more complicated. The wave crests and troughs up­ structures in a specific simulation case, respectively.
stream of the valve significantly increase, and the wave crests down­
stream of the valve obviously decreases. 3.1.1. First simulation group
This section investigates the effects of valve height ratio, valve length In the first group, the valve height ratio between valves 1 and 2
ratio, and valve spacing on the wave forces subjected by pipe-valve varies from 1/6 to 1. Table 1 lists the six configurations of the
structures. In the following numerical calculations, the wave height geometrical model. Fig. 11 shows the horizontal and vertical forces
was set to 0.06 m, the wave period was set to 1.5 s, thus the subsea pipe- acting on the pipe, valves 1 and 2 at different valve height ratios. Fig. 11
valve structures can be completely submerged during the wave cycle, (a) shows that the change in the valve height ratio had little effect on the
wave forces subjected by the pipe, which can be nearly neglected.

Fig. 8. Wave force time history of valves.

7
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 9. Wave elevation time history at different wave gauges.

Fig. 10. Schematic of the subsea pipe-valve structure model of the two valves.

Additionally, the vertical force acting on the pipe is greater than the
Table 1
horizontal force and plays a significant role in the wave forces acting on
Detailed structural dimensions at different valve height ratios (first group).
the completely submerged pipe.
Items Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Fig. 11(b) shows that the horizontal force acting on valve 1 changed
LPipe (m) 1.5 significantly as the height ratio increased. Fh+ increased as the height
LValve1 (m) 0.06 ratio increased due to the increased wave reflection effect caused by the
LValve2 (m) 0.06
D12 (m) 0.06
increased height of valve 1. The incident wave crest upstream of valve 1
HPipe (m) 0.06 gradually increased, thus increasing Fh+ . Meanwhile, the incident wave
HValve1 (m) 0.01 0.02 0.03 0.04 0.05 0.06 trough upstream of valve 1 gradually decreased, thus gradually
HValve2 (m) 0.06 increasing the free surface height difference between the left and right
HValve1/HValve2 1/6 2/6 3/6 4/6 5/6 6/6
sides of valve 1. Thus, the horizontal pressure difference formed along
(a) Pipe (b) Valve 1 (c) Valve 2. the -x-axis direction gradually increased when the wave trough passed
through valve 1, resulting in a decrease in Fh− . The relationship between

Fig. 11. Wave forces acting on each structure at different valve height ratios (the height of valve 2 remained unchanged).

8
Y. Du et al. Ocean Engineering 235 (2021) 109439

the gradual decrease in Fhmean and the simultaneous increase in the height primarily because of the wave reflection effect produced by valve 2. The
of valve 1 was nearly linear. Fig. 11(b) shows that Fv− and Fvmean first reflected wave propagated to the left, and the reflection effect increased
decreased then increased; the minimum values of Fv− and Fvmean appeared as the height of valve 2 increased. Thus, the wave force acting on valve 1
at a valve height ratio of 1/2, primarily because the reflected waves was partially offset by the reflected wave, resulting in a decrease in Fh+ .
produced by the two valves were coupled above valve 1. A super­ Fig. 16 shows that the effect of the reflected wave on valve 1 increased as
imposed wave with a large amplitude was formed, and a vertical the height of valve 2 increased when the overtopping wave impacted
slamming force was produced on valve 1 along the -z-axis direction valve 2. Additionally, the free surface height difference between the left
(Fig. 12). However, as the height of valve 1 continued increasing, the and right sides of valve 1 gradually increased as the height of valve 2
incident wave propagation to the right was gradually blocked, resulting increased. Therefore, when the wave trough passed through valve 1, the
in an increase in Fv− and Fvmean . horizontal pressure difference along the -x-axis direction gradually
Fig. 11(c) shows that Fh+ acting on valve 2 increased then decreased as increased, resulting in a decrease in Fh− . Additionally, the relationship
the valve height ratio increased. Due to the wave reflection effect generated between the gradual decrease in Fhmean and the simultaneous increase in
by valve 1, the wave crest upstream of valve 2 increased as the height of the height of valve 2 was nearly linear. Fig. 15(b) also shows that Fv+ and
valve 1 increased, resulting in an increase in Fh+ . However, for ratios greater Fvmean acting on valve 1 changed little as the height of valve 2 increased.
than 1/2, the forward propagation of the incident wave was gradually Fv− acting on valve 1 was primarily derived from the overtopping wave
blocked, and the amount of water acting on valve 2 was reduced, resulting slamming force caused by the wave run-up over valve 1, which gradu­
in a gradual decrease in Fh+ acting on valve 2. Additionally, Fhmean and Fh− ally decreased as the height ratio decreased from 6 to 6/4. Then, the
acting on valve 2 increased as the height of valve 1 increased, because as change in Fv− was irregular but relatively small, and the minimum value
the valve 1 height increased, the horizontal pressure difference along the x- of Fv− appeared at a height ratio of 6/4. The main reason is that the
axis direction will increase when the wave trough passed through valve 2. gradual increase in the height of valve 2 had a certain effect on the
Fig. 11(c) shows that Fv+ and Fvmean changed relatively little compared with motion behaviour of the overtopping wave, thereby affecting the over­
Fv− acting on valve 2 as the height of valve 1 increased; further, the mini­ topping wave force acting on valve 1.
mum value of Fv− appeared at a valve height ratio of 1/2. For relatively Fig. 15(c) shows that as the height of valve 2 increased, Fh− acting on
small valve height ratios, the vertical force acting on valve 2 was primarily valve 2 remained basically unchanged. However, Fh+ and Fhmean increased
derived from the overtopping wave formed by the wave run-up over it significantly, and the maximum value of Fh+ acting on valve 2 appeared
(Fig. 13(a)). Additionally, the wave reflection effect generated by valve 1 at a height ratio of 6/4. The horizontal force acting on valve 2 along the
increased the incident wave crest upstream of valve 2. For relatively large x-axis direction was primarily derived from the overtopping wave
height ratios, valve 1 blocks the incident wave from propagating to the formed by wave run-up over valve 1, which basically propagated to the
right. At this time, Fig. 13(b) shows that the vertical force acting on valve 2 right in the form of horizontal projectile motion (Fig. 17). The actual
was primarily derived from the overtopping wave formed by wave run-up force area of valve 2 in the horizontal direction increased as its height
⃒ ⃒
over valve 1, thus resulting in a decrease in ⃒Fv− ⃒. Fig. 14 shows the wave increased, thus resulting in an increasing Fh+ . Comparing Figs. 13(b) and
force time history of the subsea pipe-valve structure at valve height ratios of 17 shows that the motion track of the overtopping wave was signifi­
1/6 and 1. The nonlinearity of the wave force acting on the pipe is weaker cantly different under different valve 2 heights; thus, the actual force
than that of the partially submerged valves, primarily because the pipe is area of valve 2 will be affected. The vertical force acting on valve 2 was
completely submerged below the still water level, nonlinear phenomena (e. primarily derived from the overtopping wave generated by valve 1,
g., wave reflection, wave blockage, and wave run-up) have little effect on which will produce the slamming force on valve 2 along the -z direction.
the pipe. Additionally, there are non-zero initial vertical forces acting on The slamming force is proportional to the height difference between the
⃒ ⃒
each structure, primarily because when each structure is analysed sepa­ two valves; thus, the values of ⃒Fv− ⃒ relatively large when the top of valve
rately, it will be subjected by the vertical Stokes-II wave pressure. 2 does not exceed the still water level (Fig. 15(c)). The minimum value
of Fv− appeared at a height ratio of 6/2, then Fv− increased as the height of
3.1.2. Second simulation group valve 2 increased due to the decreased height difference.
In the second group, the valve height ratios between valves 1 and 2
vary from 6 to 1. Table 2 lists six configurations of the geometrical
3.2. Effect of valve length ratio on wave forces acting on the subsea pipe-
model. Fig. 15 shows the horizontal and vertical forces acting on each
valve structure
structure at different height ratios. Unlike the two valves, the pipe was
always completely submerged, so the nonlinearity of the wave forces
The effect of the valve length ratio on wave forces subjected by the
acting on it was relatively weak. Additionally, Fig. 15(a) shows that the
subsea pipe-valve structure was studied by setting two valves with
increase in the height of valve 2 had little effect on the wave forces
specified lengths on a pipe. The numerical simulations were divided into
subjected by the pipe, which can basically be neglected.
two groups. In the first group of simulations, the length of valve 2
Fig. 15(b) shows that Fh+ acting on valve 1 changed little when the
remained at 0.06 m, and the length of valve 1 increased from 0.01 m to
height ratio decreased from 6 to 2 then decreased significantly,
0.06 m at 0.01 m increments. In the second group of simulations, the

Fig. 12. Wave profile around the subsea pipe-valve structure under case 3 within one wave period.

9
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 13. Overtopping wave acting on valve 2 at different valve height ratios.

Fig. 14. Wave force time history of each structure at valve height ratios of 1/6 and 1.

length of valve 1 remained at 0.06 m, and the length of valve 2 increased


Table 2
from 0.01 m to 0.06 m at 0.01 m increments.
Detailed structural dimensions at different valve height ratios (second group).
Items Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 3.2.1. First simulation group
LPipe (m) 1.5 Table 3 lists the detailed dimensions of the subsea pipe-valve struc­
LValve1 (m) 0.06 ture model in the first group; the valve length ratio between valves 1 and
LValve2 (m) 0.06
2 varies from 1/6 to 1. Fig. 18(a) shows that the variation in the length
D12 (m) 0.06
HPipe (m) 0.06
ratio has little effect on the wave forces acting on the pipe in the hori­
HValve1 (m) 0.06 zontal and vertical directions; the corresponding reasons for the nu­
HValve2 (m) 0.01 0.02 0.03 0.04 0.05 0.06 merical results are the same as those in section 3.1.
HValve1/HValve2 6/1 6/2 6/3 6/4 6/5 6/6 The horizontal force acting on valve 1 was primarily derived from
the upstream incident wave, and changes in the length of valve 1 had
little effect on the wave shape. Thus, Fig. 18(b) shows that Fh+ and Fhmean
acting on valve 1 were basically independent of the increased length

10
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 15. Wave forces acting on each structure at different valve height ratios (the height of valve 1 remained unchanged).

Fig. 16. Velocity vector diagram of the overtopping wave impacts on valve 2 at different height ratios.

Fig. 17. Wave profile diagram of the overtopping wave impacts on valve 2 for a valve height ratio of 6/4.

ratio. Additionally, the relationship between the decrease in Fv+ , Fv− , and remained unchanged as the length of valve 1 increased. Fh+ acting on
Fvmean acting on valve 1 and the simultaneous increase in the length ratio valve 2 was primarily derived from the overtopping wave formed by the
was approximately linear. The obvious linear relationship was primarily wave run-up over valve 1; additionally, the distance between the valve 2
due to the force caused by the wave pressure during individual structure and left side of valve 1 gradually increased. Thus, for relatively short
analysis. Additionally, the difference between Fvmean and Fv− increased as valve 1 lengths, the overtopping wave had a greater impact on valve 2,
the length of valve 1 increased. The vertical force acting on valve 1 was resulting in larger Fh+ ; the maximum Fh+ value appeared at a ratio of 2/6.
primarily derived from the overtopping wave slamming force formed by As the length valve 1 continued increasing, the effect of the overtopping
the wave run-up over it, and the slamming force was proportional to the wave on valve 2 weakened; thus, Fh+ acting on valve 2 decreased. Fig. 18
slamming area projected on valve 1 (Park et al., 2013). (c) shows that as the length ratio increased, Fv+ and Fvmean remained
Fig. 18(c) shows that Fh− and Fhmean acting on valve 2 basically almost constant; Fv− underwent a relatively large change. The vertical

11
Y. Du et al. Ocean Engineering 235 (2021) 109439

Table 3 a length ratio of 1, might be because of the left-propagating reflected


Detailed structural dimensions at different valve length ratios (first group). wave generated by valve 2. In the other five cases, Fh− remained almost
Items Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 constant. Fig. 20(b) also shows that as the valve length ratio decreases,
Fv+ and Fvmean acting on valve 1 basically remained unchanged; however,
LPipe 1.5
LValve1 0.01 0.02 0.03 0.04 0.05 0.06 the change in Fv− was relatively large. The overtopping wave generated
LValve2 0.06 by the wave run-up over valve 1 played a significant role in the vertical
LValve1/LValve2 1/6 2/6 3/6 4/6 5/6 6/6 force acting on it. Fig. 21 shows that the change in valve 2 length
D12 0.06
affected the overtopping wave motion behaviour, thus affecting Fv−
HPipe 0.06
HValve1 0.06 acting on valve 1 within a certain range.
HValve2 0.06 The horizontal force acting on valve 2 was primarily derived from
the right-propagating overtopping wave generated by valve 1, and the
change in the length of valve 2 had little effect on its horizontal force
force acting on valve 2 was primarily derived from the overtopping wave
area. The motion behaviour of the overtopping wave changed as the
generated by valve 1. The overtopping wave propagated to the right in
length of valve 2 increased, resulting in slight changes in Fh+ and Fh−
the form of horizontal projectile motion, producing a slamming force
acting on valve 2 (Fig. 20(c)). Fhmean acting on valve 2 remained almost
acting on valve 2 along the -z direction. However, the change in the
constant, which indicating that the change in the length of valve 2 had
length of valve 1 affected the motion behaviour of the overtopping wave
little effect on the overall change in the horizontal force acting on it.
(Fig. 19), thereby affecting the Fv− acting on valve 2 within a certain
Fig. 20(c) also shows that the overall trend of Fv+ , Fv− , and Fvmean acting on
range.
valve 2 decreased as the length of valve 2 increased. The relationship
between the forces and the length ratio was approximately linear,
3.2.2. Second simulation group
In the second group, the length of valve 1 remained at 0.06 m, and the
length of valve 2 increased from 0.01 m to 0.06 m at 0.01 m increments. Table 4
Table 4 lists the detailed dimensions of the subsea pipe-valve model. Detailed structural dimensions at different valve length ratios (second group).
Similar to the above-mentioned simulation results, the horizontal and
Items Case 1 Case 2 Case 3 Case 4 Case 5 Case 6
vertical forces acting on the pipe were basically independent of the struc­
tural changes (Fig. 20(a)); and the vertical force acting on the pipe was LPipe 1.5
LValve1 0.06
greater than the horizontal force.
LValve2 0.01 0.02 0.03 0.04 0.05 0.06
The horizontal force acting on valve 1 was primarily derived from LValve1/LValve2 6/1 6/2 6/3 6/4 6/5 6/6
the incident wave. Because the dimension remained unchanged, valve 1 D12 0.06
had little effect on the upstream wave crest and trough. Fig. 20(b) shows HPipe 0.06
HValve1 0.06
that as the valve length ratio decreased, Fh+ acting on valve 1 changed
HValve2 0.06
slightly; while Fhmean remained almost constant. Fh− suddenly decreased at

Fig. 18. Wave forces acting on each structure at different valve length ratios (the length of valve 2 remained unchanged).

Fig. 19. Streamline diagram of overtopping wave impacts on valve 2 (the length of valve 2 remained unchanged).

12
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 20. Wave forces acting on each structure at different valve length ratios (the length of valve 1 remained unchanged).

Fig. 21. Streamline diagram of overtopping wave impacts on valve 2 (the length of valve 1 remained unchanged).

primarily due to the force caused by the wave pressure when valve 2 was reflection effect produced by valve 2 under small spacing; additionally, the
⃒ ⃒
analysed individually. Additionally, the difference between Fvmean and Fv− reflected wave produced a large ⃒Fh− ⃒ acting on valve 1 along the -x-axis
increased as the length of valve 2 increased. Fv− acting on valve 2 was direction. As the valve spacing continued increasing, the reflection effect
primarily derived from the overtopping wave slamming force generated weakened; thus Fh− was basically independent of the spacing at spacings
by valve 1 (Fig. 21); the slamming force is proportional to the force area greater than 0.04 m. Fig. 23(b) also shows that valve 1 experienced the
⃒ ⃒
of valve 2 in the vertical direction. Moreover, the right-propagating maximum ⃒Fv− ⃒ value at a valve spacing of 0.02 m, primarily because the
overtopping wave was affected by the change in the length of valve 2. effect of the reflected wave generated by valve 2 on valve 1 was relatively
strong when spacing was small; thus, a superimposed wave with a larger
3.3. Effect of valve spacing on wave forces acting on the subsea pipe-valve amplitude was formed above valve 1. Unlike Fh− , Fv− acting on valve 1
structure experienced a relatively large change as the spacing increased for spacing
greater than 0.04 m. As the valve spacing increased, the reflection effect of
The effect of the valve spacing on wave forces subjected by the valve 2 on valve 1 weakened but affected the motion behaviour of the
subsea pipe-valve structure was studied by setting three valves with overtopping wave, thereby the Fv− acting on valve 1 was affected within a
heights and lengths of 0.06 m on a pipe (Fig. 22). The spacing between certain range. The results obtained indicated that for valve 1, the sensitivity
valve 1 and valve 2 (D12) was equal to the spacing between valve 2 and of horizontal force to the variation of valve spacing was less than that of
valve 3 (D23), increased from 0.02 to 0.1 m at 0.02 m increments. We vertical force.
used the presented numerical method to calculate the wave forces acting Fig. 23(c) shows that valve 2 experienced maximum Fh+ and minimum
on each structure at different spacings. Fig. 23(a) shows that the change Fh values at spacings of 0.02 and 0.1 m, respectively. Fh+ acting on valve 2
+
in valve spacing had little effect on the wave forces subjected by the was derived from the overtopping wave impact force caused by wave run-
pipe, corresponding reasons were the same as in previous sections. up over valve 1; the effect of the wave impact force on valve 2 weakened as
Fig. 23(b) shows that Fh+ acting on valve 1 was basically the same for the spacing increased. Fig. 24(a) shows that for a valve spacing of 0.02 m,
different spacings and was primarily derived from the upstream incident the overtopping wave directly impacted valve 2 in the form of horizontal
wave. However, Fh− acting on valve 1 was much smaller at a valve spacing projectile motion. As the spacing continued increasing (Figs. 24(b) and (c)),
of 0.02 m, primarily because valve 1 was subjected by the strong wave

Fig. 22. Schematic of the subsea pipe-valve structure model with three valves.

13
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 23. Wave forces acting on each structure at different valve spacings.

the overtopping wave impact point was located in the groove between spacing from 0.02 to 0.06 m, Fv− acting on valve 3 was basically inde­
⃒ ⃒
valves 1 and 2. Additionally, ⃒Fh− ⃒ acting on valve 2 was relatively large at pendent of the valve spacing; as the spacing continued increasing, Fv−
small valve spacings of 0.02 m and 0.04 m, primarily due to the wave increased obviously. The vertical force acting on valve 3 was primarily
reflection effect generated by valve 3. The effect of the left-propagating derived from the overtopping wave slamming force generated by the
reflected wave on valve 2 weakened as the spacing increased. The verti­ other two valves. Therefore, for spacing greater than one valve length,
cal force acting on valve 2 changed less than the horizontal force as the the effect of the overtopping wave on valve 3 in the vertical direction
spacing increased; Fv+ and Fvmean were basically independent of the valve gradually weakened. The results obtained indicated that for valve 3, the
⃒ ⃒
spacing. The ⃒Fv− ⃒ acting on valve 2 was relatively small at a spacing of 0.1 sensitivity of vertical force to the variation of valve spacing was less than
m but remained largely unchanged in the other cases. The results obtained that of horizontal force.
indicated that for valve 2, the sensitivity of vertical force to the variation of
valve spacing was less than that of horizontal force. 4. Conclusions
Fig. 23(d) shows that Fh+ acting on valve 3 increased then decreased
We performed numerical simulations based on the OpenFOAM
as the valve spacing increased; the maximum value of Fh+ appeared at a
software package to study the interaction between nonlinear waves and
valve spacing of 0.04 m. The horizontal wave force acting on valve 3 was
partially submerged subsea pipe-valve structures. The subsea pipe-valve
primarily derived from the overtopping wave generated by the upstream
as a kind of multi-scale structure has more complicated hydrodynamic
valves. The overtopping wave propagated to the right and directly
performance under nonlinear incident waves than other simple struc­
impacted valve 3 at relatively small valve spacings, thus forming a
tures such as cylinders and boxes. We performed a series of numerical
relatively large Fh+ . As the spacing continued increasing, a large number
calculations of the interaction between Stokes-II waves and a semi-
of overtopping waves fell into the grooves between valves (Figs. 24(b) immersed horizontal cylinder to study the mesh and time convergence
and (c)); thus, Fh+ acting on valve 3 gradually decreased. Fig. 23(d) and to validate the accuracy of the presented numerical model. Further,
shows that the vertical force acting on valve 3 changed slightly we compared the modelling results with the experimental data, pub­
compared to the horizontal force as the spacing increased. For valve lished numerical results, and theoretical results. The calculations are

14
Y. Du et al. Ocean Engineering 235 (2021) 109439

Fig. 24. Wave profile diagram of the overtopping wave propagates to the right at different valve spacings.

mesh and time convergent, and the presented numerical method has the wave force acting on the pipe was relatively small compared
good accuracy and high efficiency. The conclusions of the research are with that of the partially submerged valves. Additionally,
summarised as follows: changes in the valve structure and valve distribution had little
effect on wave forces subjected by the pipe. The modelling results
(1) Compared with an independent valve structure, the wave forces showed that the wave force acting on the partially submerged
acting on a subsea pipe-valve structure had strong nonlinearity, valves was strongly nonlinear, which indirectly indicated that the
and the valve of the pipe-valve structure was subjected to larger wave force cannot be accurately predicted by theoretical
horizontal and vertical forces. Additionally, for a subsea pipe- equations.
valve structure, the nonlinearity of incident waves near the (3) Numerical calculations of the wave force acting on the valves
valve was significantly enhanced, the motion track of incident under different valve height and length ratios revealed that the
waves became obviously complicated. horizontal force acting on valve 1 was primarily derived from the
(2) We studied the effect of changes in valve structure and valve incident wave and from the reflected wave generated by valve 2.
distribution on wave forces acting on the subsea pipe-valve The horizontal force acting on valve 2 was primarily derived from
structures by setting different valve height ratios, valve length the incident wave upstream and from the overtopping wave
ratios, and valve spacings. The interaction between nonlinear formed by the wave run-up over valve 1. The vertical wave force
waves and subsea pipe-valve structures produces strong acting on valve 1 was primarily derived from the overtopping
nonlinear phenomena near the free surface, such as wave wave that it generated and from the reflected wave generated by
reflection, wave blockage, and wave run-up. Because the pipe valve 2. The vertical wave force acting on valve 2 was primarily
was always in a completely submerged state, the nonlinearity of derived from the overtopping wave formed by the wave run-up

15
Y. Du et al. Ocean Engineering 235 (2021) 109439

over valves 1 and 2. The overtopping wave propagated to the Fang, Q., Hong, R., Guo, A., Stansby, P.K., Li, H., 2018. Analysis of hydrodynamic forces
acting on submerged decks of coastal bridges under oblique wave action based on
right in the form of horizontal projectile motion; the motion
potential flow theory. Ocean Eng. 169, 242–252.
behaviour of the overtopping wave was affected by valve struc­ Finnegan, W., Goggins, J., 2015. Linear irregular wave generation in a numerical wave
ture changes. tank. Appl. Ocean Res. 52, 188–200.
(4) Numerical calculations of the wave force acting on the valves Gao, J., He, Z., Zang, J., Chen, Q., Ding, H., Wang, G., 2020. Numerical investigations of
wave loads on fixed box in front of vertical wall with a narrow gap under wave
under different valve spacings revealed that, for valves 2 and 3, actions. Ocean Eng. 206.
the spacing had a greater influence on the horizontal force acting Ghezelbashan, A., D’Mello, C., 2017. Wave loading on a flexible offshore structure. In:
on the valves than the vertical force. But for valve 1, the result The 27th International Ocean and Polar Engineering Conference. San Francisco,
California, USA, ISOPE-I-17-599.
was the opposite of the above. The maximum value of Fh+ and the Hong, C., Estefen, S.F., Wang, Y., Lourenço, M.I., 2018. An integrated optimization
minimum values of Fh− and Fv− acting on the valves usually model for the layout design of a subsea production system. Appl. Ocean Res. 77,
1–13.
appeared under relatively small valve spacing. The horizontal
Hu, Z.Z., Greaves, D., Raby, A., 2016. Numerical wave tank study of extreme waves and
and vertical forces acting on the valves were primarily derived wave-structure interaction using OpenFoam®. Ocean Eng. 126, 329–342.
from the incident, reflected, and overtopping waves. The latter Huang, B., Zhu, B., Cui, S., Duan, L., Zhang, J., 2018. Experimental and numerical
modelling of wave forces on coastal bridge superstructures with box girders, Part I:
two types of waves were more sensitive to spacing changes, and
regular waves. Ocean Eng. 149, 53–77.
the change in valve spacing affected the motion behaviour of the Islam, H., Mohapatra, S.C., Gadelho, J., Guedes Soares, C., 2019. OpenFOAM analysis of
overtopping wave. the wave radiation by a box-type floating structure. Ocean Eng. 193.
Jacobsen, N.G., Fuhrman, D.R., FredsøE, Jørgen, 2012. A wave generation toolbox for the
open-source CFD library: OpenFoam. Int. J. Numer. Methods Fluid. 70 (9),
However, in practical ocean engineering, a group of pipe-valve 1073–1088.
structures is generally surrounded by other groups of pipe-valve struc­ Jalón, M.L., Lira-Loarca, A., Baquerizo, A., Losada, M.Á., 2019. An analytical model for
tures or the external support frames of subsea equipment. These pipe- oblique wave interaction with a partially reflective harbor structure. Coast. Eng.
143, 38–49.
valve structures are often arranged in parallel inside the subsea equip­ Jamaludin, M.A., 2019. Cost-effective application of subsea production system as field
ment. Therefore, studying the hydrodynamic performances of the subsea development concept in shallow water environment. In: Abu Dhabi International
pipe-valve structures using 2D numerical models has limitations, Petroleum Exhibition & Conference. Abu Dhabi, UAE, SPE-197604-MS.
Lamberti, A., Martinelli, L., Gabriella Gaeta, M., Tirindelli, M., Alderson, J., 2011.
ignored the influence of the surrounding structures on the pipe-valve. Experimental spatial correlation of wave loads on front decks. J. Hydraul. Res. 49
Future work will investigate the interaction between nonlinear waves (Suppl. 1), 81–90.
and pipe-valve structures within 3D subsea equipment. Le Méhauté, B., 1976. An Introduction to Hydrodynamics and Water Waves. Springer,
ISBN 0-387-07232-2.
Li, A.-j., Liu, Y., Liu, X., Zhao, Y., 2020a. Analytical and experimental studies on water
Declaration of competing interest wave interaction with a submerged perforated quarter-circular caisson breakwater.
Appl. Ocean Res. 101.
Lindseth, S., Røsby, E., Vist, B., Aarnes, K.A., 2019. Aasta Hansteen Subsea Production
The authors declare that they have no known competing financial
System for Deep Water and Harsh Environment. Offshore Technology Conference.
interests or personal relationships that could have appeared to influence Houston, Texas, OTC-29561-MS.
the work reported in this paper. Marques Machado, F.M., Gameiro Lopes, A.M., Ferreira, A.D., 2018. Numerical
simulation of regular waves: optimization of a numerical wave tank. Ocean Eng. 170,
89–99.
Acknowledgements Martin, P.A., Dixon, A.G., 1983. The Scattering of regular surface waves by a fixed,
halfimmersed, circular cylinder. Appl. Ocean Res. 5 (1).
The authors are grateful for the financial support from the National Martínez-Ferrer, P.J., Qian, L., Ma, Z., Causon, D.M., Mingham, C.G., 2018. Improved
numerical wave generation for modelling ocean and coastal engineering problems.
Key Research and Development Program of China (Grant No. Ocean Eng. 152, 257–272.
2016YFC0303701), the National Natural Science Foundation of China Mason, P.T., Upchurch, J.L., 1996. Seastar: Subsea Cluster Manifold System Design and
(Grant No. 51879271), the 111 Project (B18054), the Brazilian Research Installation. Offshore Technology Conference. Houston, Texas, OTC-8130-MS.
Miquel, A., Kamath, A., Alagan Chella, M., Archetti, R., Bihs, H., 2018. Analysis of
Council – CNPq (Grant No. 305657/2017-8), and the Carlos Chagas different methods for wave generation and absorption in a CFD-based numerical
Filho Foundation - FAPERJ (Grant No. E-26/202.600/2019). wave tank. J. Mar. Sci. Eng. 6 (2).
Park, Y.-S., Kim, W.-J., Nam, B.-W., 2013. CFD simulation of hydrodynamic forces acting
on subsea manifold templates at wave zone. In: The Twenty-Third International
References
Offshore and Polar Engineering Conference. Anchorage, Alaska. ISOPE-I-13-347.
Rodríguez, M., Spinneken, J., 2016. A laboratory study on the loading and motion of a
Aggarwal, A., Bihs, H., Shirinov, S., Myrhaug, D., 2019. Estimation of breaking wave heaving box. J. Fluid Struct. 64, 107–126.
properties and their interaction with a jacket structure. J. Fluid Struct. 91. Rusche, Henrik, 2003. Computational Fluid Dynamics of Dispersed Two-phase Flows at
Anbarsooz, M., Passandideh-Fard, M., Moghiman, M., 2013. Fully nonlinear viscous High Phase Fractions.
wave generation in numerical wave tanks. Ocean Eng. 59, 73–85. Sayeed, T., Colbourne, B., Molyneux, D., Akinturk, A., 2017. Experimental and numerical
Bahena-Jimenez, S., Bautista, E., Méndez, F., Quesada-Torres, A., 2020. Wave reflection investigation of wave forces on partially submerged bodies in close proximity to a
by a submerged cycloidal breakwater in presence of a beach with different depth fixed structure. Ocean Eng. 132, 70–91.
profiles. Wave Motion 98 (1). Tan Loh, T., Pizer, D., Simmonds, D., Kyte, A., Greaves, D., 2018. Simulation and analysis
Bai, Y., Bai, Q., 2019. Subsea manifolds. Subsea Eng. Handb. 517–571. of wave-structure interactions for a semi-immersed horizontal cylinder. Ocean Eng.
Bernt, T., 2004. Subsea Facilities. Offshore Technology Conference. Houston, Texas, 147, 676–689.
OTC-16553-MS. Veritas, D.N., 2011. Modelling and Analysis of Marine Operations. DNV-RP-H103, Det
Buccino, M., Martinelli, L., 2020. Interaction between waves and maritime structures. Norske Veritas, Høvik, Norway.
Water 12 (12). Wu, Y.-T., Hsiao, S.-C., 2018. Generation of stable and accurate solitary waves in a
Cao, Y., Hu, X., Zhang, S., Xu, S., Lee, J., Yu, J., 2016. Design of a novel installation viscous numerical wave tank. Ocean Eng. 167, 102–113.
device for a subsea production system. Appl. Ocean Res. 59, 24–37. Yan, B., Luo, M., Bai, W., 2019. An experimental and numerical study of plunging wave
Dao, M.H., Chew, L.W., Zhang, Y., 2018. Modelling physical wave tank with flap paddle impact on a box-shape structure. Mar. Struct. 66, 272–287.
and porous beach in OpenFOAM. Ocean Eng. 154, 204–215.
Dixon, A.G., Greated, C.A., Salter, S.H., 1979. Wave forces on partially submerged
cylinders. J. Waterway 105 (4), 421–438.

16

You might also like